and ligand-gated ion channels as common mechan

6 downloads 0 Views 319KB Size Report
ingestion of house dust, but also through intake of vegetable and animal ...... Mundy WR, Freudenrich TM, Crofton KM, DeVito MJ (2004) ... mance at school age.
Environ Sci Pollut Res DOI 10.1007/s11356-013-1759-x

PCB MIXTURES IN A COMPLEX WORLD

Modulation of cell viability, oxidative stress, calcium homeostasis, and voltage- and ligand-gated ion channels as common mechanisms of action of (mixtures of) non-dioxinlike polychlorinated biphenyls and polybrominated diphenyl ethers Remco H. S. Westerink

Received: 2 November 2012 / Accepted: 22 April 2013 # Springer-Verlag Berlin Heidelberg 2013

Abstract Non-dioxin-like polychlorinated biphenyls (NDLPCBs) and polybrominated diphenyl ethers (PBDEs) are environmental pollutants that exert neurodevelopmental and neurobehavioral effects in vivo in humans and animals. Acute in vitro neurotoxic effects include changes in cell viability, oxidative stress, and basal intracellular calcium levels. Though these acute cellular effects could partly explain the observed in vivo effects, other mechanisms, such as effects on calcium influx and neurotransmitter receptor function, likely contribute to the disturbance in neurotransmission. This concise review combines in vitro data on cell viability, oxidative stress and basal calcium levels with recent data that clearly demonstrate that (hydroxylated) PCBs and (hydroxylated) PBDEs can exert acute effects on voltage-gated Ca2+ channels as well as on excitatory and inhibitory neurotransmitter receptors in vitro. These novel mechanisms of action are shared by NDL-PCBs, OH-PBDEs, and some other persistent organic pollutants, such as tetrabromobisphenol-A, and could have profound effects on neurodevelopment, neurotransmission, and neurobehavior in vivo. Keywords Non-Dioxin-Like Polychlorinated Biphenyls . Polybrominated Diphenyl Ethers . Mixture toxicity . Neurotoxicity . Cell viability . Oxidative stress . Calcium

Responsible editor: Henner Hollert R. H. S. Westerink (*) Neurotoxicology Research Group, Toxicology Division, Institute for Risk Assessment Sciences (IRAS), Faculty of Veterinary Medicine, Utrecht University, P.O. Box 80.177, 3508 TD, Utrecht, The Netherlands e-mail: [email protected]

homeostasis . Voltage-gated calcium channels . GABAA receptors . Nicotinic acetylcholine receptors

Introduction Polychlorinated biphenyls (PCBs) and polybrominated diphenyl ethers (PBDEs) are persistent organic pollutants (POPs). PCBs and PBDEs have a structural similarity and for both groups of compounds, theoretically 209 different possible congeners exist. PCBs have been used in numerous industrial and commercial applications, whereas PBDEs have been mainly used as flame retardants. Worldwide, more than 1.5 million tons of PCBs were produced until their production, commercialization, and use were largely prohibited in the 1970s and 1980s (Breivik et al. 2002). Despite this ban, PCBs are still present in the environment and in biota, mainly because of improper disposal in combination with their lipophilicity and biopersistence (ATSDR 2000; Consonni et al. 2012). Comparable to PCBs, PBDEs are globally dispersed throughout the environment (reviewed by Law et al. 2008). Since 2003, voluntary phase out of manufacturing PBDEs and use of substitutes have been taking place in North America and Europe, resulting in declining commercial PBDE production. Nonetheless, the levels of some PBDEs in the environment may still be increasing (Darnerud et al. 2001; Hites 2004), with human and environmental levels of PBDEs being approximately one order of magnitude higher in North America than in Europe and Asia (reviewed by Frederiksen et al. 2009). Human exposure to PCBs occurs mainly via food intake (EFSA 2005), although inhalation of (indoor) air and house

Environ Sci Pollut Res

dust also provides significant exposure, mainly to lowerchlorinated congeners (Broding et al. 2007; Hu et al. 2012). Human exposure to PBDEs occurs in particular via air and ingestion of house dust, but also through intake of vegetable and animal products (reviewed by Frederiksen et al. 2009) and in occupational settings (reviewed by Carpenter 2006; Frederiksen et al. 2009). Children are exposed to larger amounts of these POPs than adults (Fischer et al. 2006; Toms et al. 2009), mainly due to their relatively high food intake, child-specific hand-to-mouth behavior, and larger ingestion of house dust. An additional source of exposure for young children is breast milk, in which especially lowerbrominated PBDEs have been detected (reviewed by Frederiksen et al. 2009). Regardless of the route of exposure, levels of individual PCBs and PBDEs in human (cord) blood have been shown to be in the subnanomolar to lower nanomolar range, whereas the sum of the levels of the most common PCBs and PBDEs in blood can add up to the higher nanomolar range, especially in occupationally exposed people (Gabrio et al. 2000; Petrik et al. 2006; reviewed in Dingemans et al. 2011). PCBs are divided into coplanar dioxin-like (DL-) PCBs and non-dioxin-like (NDL-) PCBs. Contrary to DL-PCBs, NDL-PCBs have one or more chlorines in the ortho-positions, resulting in nonplanar structures. As a result, NDLPCBs have little or no affinity to the aryl hydrocarbon receptor (AhR) and display a different toxicological profile (ATSDR 2000; Van den Berg et al. 2006). Epidemiological studies indicated that perinatal exposure to NDL-PCBs can result in neurodevelopmental and neurobehavioral effects in children (for review see Faroon et al. 2001; Winneke et al. 2002; Schantz et al. 2003). More recently, adverse effects on cognitive and neurodevelopmental parameters in humans were also correlated to PBDE exposure. Motor, cognitive, and behavioral performance in 6-year-old Dutch children was correlated with maternal serum levels of PBDEs measured in the 35th week of pregnancy (Roze et al. 2009). In another study, the scores of US children 0–6 years of age on yearly tests of mental and physical development were lower among those prenatally exposed to higher concentrations of PBDEs (Herbstman et al. 2010). These findings are in line with in vivo animal studies that demonstrated that (neonatal) exposure to PCBs and PBDEs can induce a wide range of neurobehavioral effects, including changes in motor activity, spontaneous behavior, learning, memory, and attention (reviewed in Ulbrich and Stahlmann 2004; Costa and Giordano 2007; Fonnum and Mariussen 2009; Dingemans et al. 2011). To date, in vitro studies identified several mechanisms of action for NDL-PCBs and PBDEs, including alterations in cell viability, oxidative stress, calcium homeostasis, as well as on neurotransmitter receptor function (see below), that could underlie these observed in vivo effects in animals and humans.

Effects on cell viability and oxidative stress A reduction in cell viability is a clear, yet not very subtle sign of xenobiotic-induced neurotoxicity. Similarly, xenobiotic-induced increased oxidative stress, e.g., enhanced production of reactive oxygen species (ROS), is a clear indication for neurotoxicity. Both, cell viability and oxidative stress can be assessed in vitro using highthroughput biochemical assays. It is therefore not surprising that many of the early in vitro studies on PCBs, dating back to the 1970s and 1980s, and on PBDEs focused on effects on cell viability. Currently, effects of PCBs and PBDEs on cell viability have been assessed in numerous cell models, including primary cultures, and usually occur in the (tens of) micromolar range. The effects on cell viability often closely correlate with induction of oxidative stress, suggesting that induction of oxidative stress is causally related to the reduction in cell viability (see e.g., Howard et al. 2003; Huang et al. 2010). In support of this, the reduction in cell viability induced by PCBs and PBDEs could be (partially) prevented by the addition of antioxidants such as vitamin E and melatonin (see e.g., Fonnum et al. 2006; Costa and Giordano 2007; Costa et al. 2008; Fonnum and Mariussen 2009; Dingemans et al. 2011 for more extensive reviews on cell viability and oxidative stress). Despite the wealth of available studies and the relative simplicity of the abovementioned endpoints, several issues remain to be completely resolved. These issues are mainly related to bioavailability, bioaccumulation, bioactivation, and the possible occurrence of additive or synergistic effects following exposure to mixtures of POPs. Previous studies indicated that PBDEs can accumulate intracellularly due to their high lipophilicity and, as such, concentrations in the medium may be underestimating the concentration present in the cell by 1-2 orders of magnitude (Mundy et al. 2004; Kodavanti et al. 2005; Huang et al. 2010). On the other hand, inclusion of serum in the exposure medium strongly (∼5-fold) attenuated the intracellular accumulation of PBDEs, indicating reduced bioavailability due to binding of PBDEs to serum proteins (Mundy et al. 2004). Absorption, elimination, and distribution of PCBs and PBDEs have been extensively studied. Importantly, both PCBs and PBDEs can undergo oxidative metabolism in vivo resulting in numerous hydroxylated metabolites at concentrations similar to or sometimes even exceeding those of the unmetabolized parent compounds (see e.g., Costa and Giordano 2007; Fonnum and Mariussen 2009; Dingemans et al. 2011 for more extensive reviews on kinetics). These hydroxylated metabolites are reported to be more potent in reducing cell viability and increasing oxidative stress. The level of oxidative stress in e.g., cerebellar granule cells induced by hydroxylated PCBs is increased ∼10-fold compared to control (Dreiem et al. 2009), whereas parent

Environ Sci Pollut Res

compounds were reported to increase ROS levels by less than 2-fold (Mariussen et al. 2002). The degree of bioactivation with respect to cell viability and ROS production is even more pronounced for PBDEs. Hydroxylated PBDEs were shown to more effectively increase cytotoxicity in H295R human adrenocortical carcinoma cells (Cantón et al. 2006), rat PC12 cells (Dingemans et al. 2010a), and chicken DT40 cells (Ji et al. 2011), to increase oxidative stress in human hepatoma HepG2 cells (An et al. 2011) and to increase both cytotoxicity and oxidative stress in embryonic zebrafish (Usenko et al. 2012). Yet, this issue of bioactivation is far from completely resolved as any of the individual PCB or PBDE congeners can give rise to (a mixture of) multiple metabolites yielding an enormous amount of metabolites and possible exposure scenarios. Currently, a direct comparison of the parent compounds and the possible metabolites with respect to cytotoxicity and oxidative stress is often lacking and it is often unclear which of these metabolites can actually be formed in vivo. Another largely unresolved issue is the occurrence of antagonistic, additive, or synergistic effects following exposure to mixtures of environmental pollutants. For DL-PCBs, the toxic and biological effects have been primarily associated with binding and activation of the AhR transduction pathway. With respect to AhR activation, DL-PCBs have been assigned a “toxic equivalent factor” (TEF). The TEF concept relies on the relative effect potency determined for individual polychlorinated dibenzo-p-dioxins, polychlorinated dibenzofurans, and PCB congeners for inducing AhR activation relative to the reference compound 2,3,7,8-tetrachlorodibenzo-p-dioxin. When the observed toxic effects of multiple compounds are additive the total toxic equivalent (TEQ, which is defined by the sum of the products of the concentration of each compound multiplied by its TEF value) can be used to estimate the potency of the mixture (Van den Berg et al. 2006). However, despite the fact that induction of cytotoxicity and oxidative stress by PCBs and PBDEs has often been tested using commercial mixtures, such as Aroclor1254 and DE-71, it is still not clearly established whether the TEF concept, and thus additivity, applies for these endpoints. On the other hand, the TEF concept has been proposed for activation of ryanodine (Ry) receptors (RyR) on the endoplasmic reticulum (ER) by NDL-PCBs and PBDEs (Pessah et al. 2010; Kim et al. 2011) and activation of RyR correlated well with cytotoxicity (Kim et al. 2011). Establishing a TEF concept for the different neurotoxic endpoints has been hampered, at least partly, by variability in the composition of commercial mixtures (see e.g., Kodavanti et al. 2001). More importantly, only a few studies directly compare the potency of different individual congeners with the respective mixtures, often yielding contradictory results. In J774A.1 macrophage cells, PCBs 101, 153, and 180 induced a synergistic effect on cell death and apoptosis at concentrations

that were inactive for the individual congeners (Ferrante et al. 2011). Contrary, in human T47D and MDA-MB-231 breast cancer cells, weak antagonism was observed between PCB153 and PCB126 for cytotoxicity and ROS production at lower concentrations (50 % activation (at 10 μM) and >800 % potentiation (at 10 μM), whereas BDE-47 is ineffective (Hendriks et al. 2010). Interestingly, though binary mixtures of NDL-PCBs result in competitive activation and potentiation of human GABAA-Rs, a binary mixture of PCB-47 and 6-OH-PBDE-47 induced an additive activation as well as potentiation of GABAA receptors (Hendriks et al. 2010). Overall, NDL-PCBs and OH-PBDEs thus act as (partial) agonist on the postsynaptic human GABAA-Rs, thereby potentially reducing excitability of the nervous system. Noteworthy in this respect is that e.g. developmental PCB exposure was reported to increase extracellular levels of GABA, thereby potentially contributing to a reduction in neurotransmission (Boix et al. 2010). These findings thus demonstrate that GABAA and nACh receptors are affected differently by PCB-47 and 6-OHPBDE-47, with inhibitory GABAA-mediated signaling being potentiated and excitatory α4β2 nACh-mediated signaling being inhibited. Considering these opposite actions and the additive interaction of the congeners, these effects are likely to be augmented in vivo, resulting in an inhibition of neuronal excitability. This may be further enhanced by additional effects on voltage-gated ion channels, including inhibition of voltage-gated calcium channels by NDL-PCBs and (OH-) BDEs (see above: effects on calcium homeostasis). Moreover, BDE-209 (at ≥0.1 μM) was recently shown to concentration-dependently inhibit voltage-gated sodium channels (VGSCs), shift the activation and inactivation of VGSCs towards hyperpolarizing direction, slow down the recovery from inactivation and decrease the fraction of activated VGSCs (Xing et al. 2010), suggesting that BDE209

Environ Sci Pollut Res

could inhibit the generation of action potentials and thus neuronal excitability.

Evidently, NDL-PCBs and (OH-) BDEs can exert neurotoxic effects in vitro, including effects on cell viability, oxidative stress, and basal calcium levels via intracellular calcium stores. Importantly, these effects are nicely correlated, suggesting a causal relation of store-mediated increases in basal [Ca2+]i, increased oxidative stress and caspase activity with reduced cell viability (also see Fig. 1). Of concern is that the store-mediated increases in basal [Ca2+]i, already occur in the low- and even sub-micromolar level. Of similar concern, the acute effects on functional endpoints (calcium homeostasis and neurotransmitter receptor function) also occur already in the low- and even submicromolar level (see Fig. 1). Though some discrepancies exist, the effects on VGCC, VGSC, GABAA receptors, and nACh receptors suggest that (mixtures of) PCBs and PBDEs can reduce neuronal excitability. Consequently, the overall effect in vivo could be even more pronounced, though it should be noted that till date effects of PCBs or PBDEs on many neurotransmitter receptors, e.g., glutamate or glycine receptors, have not or only hardly been studied. On the other hand, the increase in basal [Ca2+]i could not only affect cell viability pathways, but can also increase neuronal excitability, thereby counteracting the suggested reduction in

neuronal excitability. As such, the overall effect on neuronal excitability remains to be determined and potentially differs between in vitro models as well as between different brain areas in vivo due to differences in relative receptor expression. For risk assessment purposes, additional information regarding bioactivation is essential since hydroxylation drastically increases the neurotoxic potency of e.g. BDE-47. Noteworthy is this respect is also the observation that binary mixtures of PCBs with (OH-) BDEs can induce additive effects, whereas additivity is apparently less likely to occur for mixtures consisting of only PCBs or only PBDEs. Hopefully, future research can fully elucidate if and under which conditions additivity applies. Notably, the adverse effects described in this review are not only shared by PCBs and PBDEs, but possibly by other POPs as well. Recently, the brominated flame retardant tetrabromobisphenol-A (TBBPA) has been shown to not only induce cell death and oxidative stress, but also to increase basal [Ca2+]i, inhibit VGCCs and α4β2 nACh-Rs, and to activate/potentiate α1β2γ2 GABAA-Rs (Hendriks et al. 2012b), just as NDL-PCBs and 6-OH-BDE-47. A thorough investigation and comparison of different abundant POPs for these specific endpoints could be a first step towards (mechanistically) investigating the effects of environmentally relevant complex mixtures. Currently, the best studied environmentally relevant combination includes methylmercury (MeHg). Mixtures of PCBs with MeHg have previously been shown to synergistically reduce brain dopamine content in rat striatal slices, while synergistically increasing extracellular

Fig. 1 Schematic representation of the common cellular actions of NDL-PCBs, OH-PBDEs, and some other POPs, such as TBBPA. Neuronal activity can be reduced by inhibitory actions on nicotinic acetylcholine receptors (nACh-R) and voltage-gated calcium channels (VGCC), resulting in reduced neurotransmitter secretion (middle). Moreover, neuronal activity and thus neurotransmitter secretion can be further reduced by inhibitory actions on voltage-gated sodium

channels (VGSC) and activating/potentiating effects on GABAA receptors. In contrast to these inhibitory effects, NDL-PCBs and OHPBDEs can induce release of calcium from intracellular stores, such as the endoplasmic reticulum and mitochondria, which could (temporarily) increase neurotransmitter secretion. More importantly, calciumstore-mediated calcium release could trigger activation of (calciumdependent, caspase-mediated) cell death pathways

Future perspectives

Environ Sci Pollut Res

dopamine levels (Bemis and Seegal 1999). Similarly, a mixture of PCB4 with MeHg synergistically increased basal calcium levels in rat cerebellar granule cells, at least when both were applied at low micromolar concentrations (Bemis and Seegal 2000). In line with these findings, co-exposure of MeHg (sub-micromolar range) with PCBs (micromolar range) resulted in additive or slightly synergistic effects on the induction of cell death via activation of calcium-regulated calpains and lysosomal cathepsins, possibly through disruption of mitochondrial function and intracellular calcium signaling in AtT20 pituitary cells (Johansson et al. 2006). Yet, this issue is also far from resolved since more recently coexposure of mouse hippocampal neuronal HT22 cells to MeHg and PCB153 or PCB126 was shown to result mainly in antagonistic effects on cell viability (Tofighi et al. 2011). Until the above matters are investigated more thoroughly, and for different exposure scenarios, risk assessment of mixtures of POPs, including PCBs and PBDEs, will be hampered by a lack of insight into the constrains of additivity and bioactivation with respect to the different neurotoxic endpoints. Acknowledgments Elsa C. Antunes Fernandes, Milou M.L. Dingemans, Hester S. Hendriks, Regina G.D.M. van Kleef, Wendy T. Langeveld, Marieke Meijer, Ad Reniers, and other members of the Neurotoxicology Research Group (IRAS) are gratefully acknowledged for their valuable scientific and technical support throughout the projects that provided the data and insight to prepare this review. I sincerely apologize to all the authors of primary literature or previous reviews that could not be included due to space limitations. This study was supported by the European Union [ATHON: FP7-FOOD-CT-2005-022923; ENFIRO: FP7-ENV-2008-1-226563; DENAMIC: grant no. FP7-ENV-2011282957] and the Faculty of Veterinary Medicine of Utrecht University.

References An J, Li S, Zhong Y, Wang Y, Zhen K, Zhang X, Wang Y, Wu M, Yu Z, Sheng G, Fu J, Huang Y (2011) The cytotoxic effects of synthetic 6-hydroxylated and 6-methoxylated polybrominated diphenyl ether 47 (BDE47). Environ Toxicol 26:591–599 Antunes Fernandes EC, Hendriks HS, van Kleef RG, van den Berg M, Westerink RH (2010a) Potentiation of the human GABA(A) receptor as a novel mode of action of lower-chlorinated nondioxin-like PCBs. Environ Sci Technol 44:2864–2869 Antunes Fernandes EC, Hendriks HS, van Kleef RG, Reniers A, Andersson PL, van den Berg M, Westerink RH (2010b) Activation and potentiation of human GABAA receptors by non-dioxin-like PCBs depends on chlorination pattern. Toxicol Sci 118:183–190 ATSDR (2000) Toxicological profile for polychlorinated biphenyls (PCBs). Agency for toxic substances and disease registry, Atlanta Bae J, Stuenkel EL, Loch-Caruso R (1999) Stimulation of oscillatory uterine contraction by the PCB mixture Aroclor 1242 may involve increased [Ca2+]i through voltage-operated calcium channels. Toxicol Appl Pharmacol 155:261–272 Bemis JC, Seegal RF (1999) Polychlorinated biphenyls and methylmercury act synergistically to reduce rat brain dopamine content in vitro. Environ Health Perspect 107:879–885

Bemis JC, Seegal RF (2000) Polychlorinated biphenyls and methylmercury alter intracellular calcium concentrations in rat cerebellar granule cells. Neurotoxicology 21:1123–1134 Berridge MJ, Bootman MD, Roderick HL (2003) Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol 4:517–529 Boix J, Cauli O, Felipo V (2010) Developmental exposure to polychlorinated biphenyls 52, 138, or 180 affects differentially learning or motor coordination in adult rats. Mechanisms involved. Neuroscience 167:994–1003 Breivik K, Sweetman A, Pacyna JM, Jones KC (2002) Towards a global historical emission inventory for selected PCB congeners-a mass balance approach. 1. Global production and consumption. Sci Total Environ 290:181–198 Broding HC, Schettgen T, Goen T, Angerer J, Drexler H (2007) Development and verification of a toxicokinetic model of polychlorinated biphenyl elimination in persons working in a contaminated building. Chemosphere 68:1427–1434 Cantón RF, Sanderson JT, Nijmeijer S, Bergman A, Letcher RJ, van den Berg M (2006) In vitro effects of brominated flame retardants and metabolites on CYP17 catalytic activity: a novel mechanism of action? Toxicol Appl Pharmacol 216:274–281 Carpenter DO (2006) Polychlorinated biphenyls (PCBs): routes of exposure and effects on human health. Rev Environ Health 21:1–23 Carrasco MA, Hidalgo C (2006) Calcium microdomains and gene expression in neurons and skeletal muscle cells. Cell Calcium 40:575–583 Catterall WA, Few AP (2008) Calcium channel regulation and presynaptic plasticity. Neuron 59:882–901 Cavazzini M, Bliss T, Emptage N (2005) Ca2+ and synaptic plasticity. Cell Calcium 38:355–367 Chishti MA, Fisher JP, Seegal RF (1996) Aroclors 1254 and 1260 reduce dopamine concentrations in rat striatal slices. Neurotoxicology 17:653–660 Coburn CG, Currás-Collazo MC, Kodavanti PRS (2008) In vitro effects of environmentally relevant polybrominated diphenyl ether (PBDE) congeners on calcium buffering mechanisms in rat brain. Neurochem Res 33:355–364 Collin T, Marty A, Llano I (2005) Presynaptic calcium stores and synaptic transmission. Curr Opin Neurobiol 15:275–281 Consonni D, Sindaco R, Bertazzi PA (2012) Blood levels of dioxins, furans, dioxin-like PCBs, and TEQs in general populations: a review, 1989–2010. Environ Int 44:151–62 Costa LG, Giordano G (2007) Developmental neurotoxicity of polybrominated diphenyl ether (PBDE) flame retardants. Neurotoxicology 28:1047–1067 Costa LG, Giordano G, Tagliaferri S, Caglieri A, Mutti A (2008) Polybrominated diphenyl ether (PBDE) flame retardants: environmental contamination, human body burden, and potential adverse health effects. Acta Biomed 79:172–183 D’Hulst C, Atack JR, Kooy RF (2009) The complexity of the GABAA receptor shapes unique pharmacological profiles. Drug Discov Today 14:866–875 Darnerud PO, Eriksen GS, Johannesson T, Larsen PB, Viluksela M (2001) Polybrominated diphenyl ethers: occurrence, dietary exposure, and toxicology. Environ Health Perspect 109(1):49–68 Dickerson SM, Guevara E, Woller MJ, Gore AC (2009) Cell death mechanisms in GT1-7 GnRH cells exposed to polychlorinated biphenyls PCB74, PCB118, and PCB153. Toxicol Appl Pharmacol 237:237–245 Dingemans MM, Ramakers GM, Gardoni F, van Kleef RG, Bergman A, Di Luca M, van den Berg M, Westerink RH, Vijverberg HP (2007) Neonatal exposure to brominated flame retardant BDE-47 reduces long-term potentiation and postsynaptic protein levels in mouse hippocampus. Environ Health Perspect 115:865–870

Environ Sci Pollut Res Dingemans MM, de Groot A, van Kleef RG, Bergman A, van den Berg M, Vijverberg HP, Westerink RH (2008) Hydroxylation increases the neurotoxic potential of BDE-47 to affect exocytosis and calcium homeostasis in PC12 cells. Environ Health Perspect 116:637–643 Dingemans MM, Heusinkveld HJ, Bergman A, van den Berg M, Westerink RH (2010a) Bromination pattern of hydroxylated metabolites of BDE-47 affects their potency to release calcium from intracellular stores in PC12 cells. Environ Health Perspect 118:519–525 Dingemans MM, van den Berg M, Bergman A, Westerink RH (2010b) Calcium-related processes involved in the inhibition of depolarization-evoked calcium increase by hydroxylated PBDEs in PC12 cells. Toxicol Sci 114:302–309 Dingemans MM, van den Berg M, Westerink RH (2011) Neurotoxicity of brominated flame retardants: (in)direct effects of parent and hydroxylated polybrominated diphenyl ethers on the (developing) nervous system. Environ Health Perspect 119:900–907 Dreiem A, Rykken S, Lehmler HJ, Robertson LW, Fonnum F (2009) Hydroxylated polychlorinated biphenyls increase reactive oxygen species formation and induce cell death in cultured cerebellar granule cells. Toxicol Appl Pharmacol 240:306–313 Dwyer JB, McQuown SC, Leslie FM (2009) The dynamic effects of nicotine on the developing brain. Pharmacol Ther 122:125–139 EFSA (2005) Opinion of the scientific panel on contaminants in the food chain on a request from the comission related to the presence of non dioxin-like polychlorinated biphenyls (PCB) in feed and food. The EFSA Journal, 1–137 Eriksson P, Fredriksson A (1996) Developmental neurotoxicity of four ortho-substituted polychlorinated biphenyls in the neonatal mouse. Environ Toxicol Pharmacol 1:155–165 Eriksson P, Viberg H, Jakobsson E, Orn U, Fredriksson A (2002) A brominated flame retardant, 2,2′,4,4′,5-pentabromodiphenyl ether: uptake, retention, and induction of neurobehavioral alterations in mice during a critical phase of neonatal brain development. Toxicol Sci 67:98–103 Fan CY, Besas J, Kodavanti PR (2010) Changes in mitogen-activated protein kinase in cerebellar granule neurons by polybrominated diphenyl ethers and polychlorinated biphenyls. Toxicol Appl Pharmacol 245:1–8 Faroon O, Jones D, de Rosa C (2001) Effects of polychlorinated biphenyls on the nervous system. Toxicol Ind Health 16:305–333 Ferrante MC, Mattace Raso G, Esposito E, Bianco G, Iacono A et al (2011) Effects of non-dioxin-like polychlorinated biphenyl congeners (PCB 101, PCB 153 and PCB 180) alone or mixed on J774A.1 macrophage cell line: modification of apoptotic pathway. Toxicol Lett 202:61–68 Fischer D, Hooper K, Athanasiadou M, Athanassiadis I, Bergman A (2006) Children show highest levels of polybrominated diphenyl ethers in a California family of four: a case study. Environ Health Perspect 114:1581–1584 Fonnum F, Mariussen E (2009) Mechanisms involved in the neurotoxic effects of environmental toxicants such as polychlorinated biphenyls and brominated flame retardants. J Neurochem 111:1327–1347 Fonnum F, Mariussen E, Reistad T (2006) Molecular mechanisms involved in the toxic effects of polychlorinated biphenyls (PCBs) and brominated flame retardants (BFRs). J Toxicol Environ Health A 69:21–35 Frederiksen M, Vorkamp K, Thomsen M, Knudsen LE (2009) Human internal and external exposure to PBDEs-a review of levels and sources. Int J Hyg Environ Health 212:109–134 Fujii S, Jia Y, Yang A, Sumikawa K (2000) Nicotine reverses GABAergic inhibition of long-term potentiation induction in the hippocampal CA1 region. Brain Res 863:259–265

Gabrio T, Piechotowski I, Wallenhorst T, Klett M, Cott L, Friebel P, Link B, Schwenk M (2000) PCB-blood levels in teachers, working in PCB-contaminated schools. Chemosphere 40:1055–1062 Gao P, He P, Wang A, Xia T, Xu B, Xu Z, Niu Q, Guo L, Chen X (2009) Influence of PCB153 on oxidative DNA damage and DNA repair-related gene expression induced by PBDE-47 in human neuroblastoma cells in vitro. Toxicol Sci 107:165–170 Garcia AG, Garcia de Diego AM, Gandia L, Borges R, Garcia-Sancho J (2006) Calcium signaling and exocytosis in adrenal chromaffin cells. Physiol Rev 86:1093–1131 Gilbert ME (2003) Perinatal exposure to polychlorinated biphenyls alters excitatory synaptic transmission and short-term plasticity in the hippocampus of the adult rat. Neurotoxicology 24:851– 860 He P, Wang AG, Xia T, Gao P, Niu Q, Guo LJ, Xu BY, Chen XM (2009) Mechanism of the neurotoxic effect of PBDE-47 and interaction of PBDE-47 and PCB153 in enhancing toxicity in SH-SY5Y cells. Neurotoxicology 30:10–15 He W, Wang A, Xia T, Gao P, Xu B, Xu Z, He P, Chen X (2010) Cytogenotoxicity induced by PBDE-47 combined with PCB153 treatment in SH-SY5Y cells. Environ Toxicol 25:564–572 Hendriks HS, Antunes Fernandes EC, Bergman A, van den Berg M, Westerink RH (2010) PCB-47, PBDE-47, and 6-OH-PBDE-47 differentially modulate human GABAA and alpha4beta2 nicotinic acetylcholine receptors. Toxicol Sci 118:635–642 Hendriks HS, van Kleef RG, Westerink RH (2012a) Modulation of human α4β2 nicotinic acetylcholine receptors by brominated and halogen-free flame retardants as a measure for in vitro neurotoxicity. Toxicol Lett 213:266–274 Hendriks HS, van Kleef RG, van den Berg M, Westerink RH (2012b) Multiple novel modes of action involved in the in vitro neurotoxic effects of tetrabromobisphenol-A. Toxicol Sci 128:235–246 Herbstman JB, Sjödin A, Kurzon M, Lederman SA, Jones RS, Rauh V et al (2010) Prenatal exposure to PBDEs and neurodevelopment. Environ Health Perspect 118:712–719 Heusinkveld HJ, Westerink RH (2011) Caveats and limitations of plate reader-based high-throughput kinetic measurements of intracellular calcium levels. Toxicol Appl Pharmacol 255:1–8 Hites RA (2004) Polybrominated diphenyl ethers in the environment and in people: a meta-analysis of concentrations. Environ Sci Technol 38:945–956 Howard AS, Fitzpatrick R, Pessah I, Kostyniak P, Lein PJ (2003) Polychlorinated biphenyls induce caspase-dependent cell death in cultured embryonic rat hippocampal but not cortical neurons via activation of the ryanodine receptor. Toxicol Appl Pharmacol 190:72–86 Hu X, Adamcakova-Dodd A, Lehmler HJ, Hu D, Hornbuckle K, Thorne PS (2012) Subchronic inhalation exposure study of an airborne polychlorinated biphenyl mixture resembling the Chicago ambient air congener profile. Environ Sci Technol 46:9653–9662 Huang SC, Giordano G, Costa LG (2010) Comparative cytotoxicity and intracellular accumulation of five polybrominated diphenyl ether congeners in mouse cerebellar granule neurons. Toxicol Sci 114:124–132 Inglefield JR, Shafer TJ (2000a) Polychlorinated biphenyl-stimulation of Ca(2+) oscillations in developing neocortical cells: a role for excitatory transmitters and L-type voltage-sensitive Ca(2+) channels. J Pharmacol Exp Ther 295:105–113 Inglefield JR, Shafer TJ (2000b) Perturbation by the PCB mixture aroclor 1254 of GABA(A) receptor-mediated calcium and chloride responses during maturation in vitro of rat neocortical cells. Toxicol Appl Pharmacol 164:184–195 Inglefield JR, Mundy WR, Shafer TJ (2001) Inositol 1,4,5-triphosphate receptor-sensitive Ca(2+) release, store-operated Ca(2+) entry, and cAMP responsive element binding protein phosphorylation

Environ Sci Pollut Res in developing cortical cells following exposure to polychlorinated biphenyls. J Pharmacol Exp Ther 297:762–773 Ji K, Choi K, Giesy JP, Musarrat J, Takeda S (2011) Genotoxicity of several polybrominated diphenyl ethers (PBDEs) and hydroxylated PBDEs, and their mechanisms of toxicity. Environ Sci Technol 45:5003–5008 Johansson C, Tofighi R, Tamm C, Goldoni M, Mutti A, Ceccatelli S (2006) Cell death mechanisms in AtT20 pituitary cells exposed to polychlorinated biphenyls (PCB 126 and PCB 153) and methylmercury. Toxicol Lett 167:183–190 Johansson N, Viberg H, Fredriksson A, Eriksson P (2008) Neonatal exposure to deca-brominated diphenyl ether (PBDE 209) causes dose–response changes in spontaneous behaviour and cholinergic susceptibility in adult mice. Neurotoxicology 29:911–919 Kim KH, Inan SY, Berman RF, Pessah IN (2009) Excitatory and inhibitory synaptic transmission is differentially influenced by two ortho-substituted polychlorinated biphenyls in the hippocampal slice preparation. Toxicol Appl Pharmacol 237:168–177 Kim KH, Bose DD, Ghogha A, Riehl J, Zhang R, Barnhart CD, Lein PJ, Pessah IN (2011) Para- and ortho-substitutions are key determinants of polybrominated diphenyl ether activity toward ryanodine receptors and neurotoxicity. Environ Health Perspect 119:519–526 Kodavanti PR, Ward TR (2005) Differential effects of commercial polybrominated diphenyl ether and polychlorinated biphenyl mixtures on intracellular signaling in rat brain in vitro. Toxicol Sci 85:952–962 Kodavanti PR, Shin DS, Tilson HA, Harry GJ (1993) Comparative effects of 2 polychlorinated biphenyl congeners on calcium homeostasis in rat cerebellar granule cells. Toxicol Appl Pharmacol 123:97–106 Kodavanti PR, Kannan N, Yamashita N, Derr-Yellin EC, Ward TR, Burgin DE, Tilson HA, Birnbaum LS (2001) Differential effects of two lots of aroclor 1254: congener-specific analysis and neurochemical end points. Environ Health Perspect 109:1153–1161 Kodavanti PR, Ward TR, Derr-Yellin EC, McKinney JD, Tilson HA (2003) Increased [3H]phorbol ester binding in rat cerebellar granule cells and inhibition of 45Ca(2+) buffering in rat cerebellum by hydroxylated polychlorinated biphenyls. Neurotoxicology 24:187–198 Kodavanti PR, Ward TR, Ludewig G, Robertson LW, Birnbaum LS (2005) Polybrominated diphenyl ether (PBDE) effects in rat neuronal cultures: 14C-PBDE accumulation, biological effects, and structure-activity relationships. Toxicol Sci 88:181–192 Langeveld WT, Meijer M, Westerink RH (2012) Differential effects of 20 non-dioxin-like PCBs on basal and depolarization-evoked intracellular calcium levels in PC12 cells. Toxicol Sci 126:487–496 Law RJ, Herzke D, Harrad S, Morris S, Bersuder P, Allchin CR (2008) Levels and trends of HBCD and BDEs in the European and Asian environments, with some information for other BFRs. Chemosphere 73:223–241 Lin CH, Lin PH (2006) Induction of ROS formation, poly(ADPribose) polymerase-1 activation, and cell death by PCB126 and PCB153 in human T47D and MDA-MB-231 breast cancer cells. Chem Biol Interact 162:181–194 Llansola M, Montoliu C, Boix J, Felipo V (2010) Polychlorinated biphenyls PCB 52, PCB 180, and PCB 138 impair the glutamate-nitric oxide-cGMP pathway in cerebellar neurons in culture by different mechanisms. Chem Res Toxicol 23:813–820 Londoño M, Shimokawa N, Miyazaki W, Iwasaki T, Koibuchi N (2010) Hydroxylated PCB induces Ca2+ oscillations and alterations of membrane potential in cultured cortical cells. J Appl Toxicol 30:334–342 Magi S, Castaldo P, Carrieri G, Scorziello A, Di Renzo G, Amoroso S (2005) Involvement of Na+-Ca2+ exchanger in intracellular Ca2+ increase and neuronal injury induced by polychlorinated

biphenyls in human neuroblastoma SH-SY5Y cells. J Pharmacol Exp Ther 315:291–296 Mariussen E, Fonnum F (2006) Neurochemical targets and behavioral effects of organohalogen compounds: an update. Crit Rev Toxicol 36:253–289 Mariussen E, Myhre O, Reistad T, Fonnum F (2002) The polychlorinated biphenyl mixture aroclor 1254 induces death of rat cerebellar granule cells: the involvement of the N-methyl-D-aspartate receptor and reactive oxygen species. Toxicol Appl Pharmacol 179:137–144 Mohler H (2007) Molecular regulation of cognitive functions and developmental plasticity: impact of GABAA receptors. J Neurochem 102:1–12 Moody WJ, Bosma MM (2005) Ion channel development, spontaneous activity, and activity-dependent development in nerve and muscle cells. Physiol Rev 85:883–941 Mundy WR, Freudenrich TM, Crofton KM, DeVito MJ (2004) Accumulation of PBDE-47 in primary cultures of rat neocortical cells. Toxicol Sci 82:164–169 Orrenius S, Nicotera P, Zhivotovsky B (2011) Cell death mechanisms and their implications in toxicology. Toxicol Sci 119:3–19 Pessah IN, Cherednichenko G, Lein PJ (2010) Minding the calcium store: ryanodine receptor activation as a convergent mechanism of PCB toxicity. Pharmacol Ther 125:260–285 Petrik J, Drobna B, Pavuk M, Jursa S, Wimmerova S, Chovancova J (2006) Serum PCBs and organochlorine pesticides in Slovakia: age, gender, and residence as determinants of organochlorine concentrations. Chemosphere 65:410–418 Reistad T, Mariussen E (2005) A commercial mixture of the brominated flame retardant pentabrominated diphenyl ether (DE-71) induces respiratory burst in human neutrophil granulocytes in vitro. Toxicol Sci 87:57–65 Reistad T, Mariussen E, Fonnum F (2006) Neurotoxicity of the pentabrominated diphenyl ether mixture, DE-71, and hexabromocyclododecane (HBCD) in rat cerebellar granule cells in vitro. Arch Toxicol 80:785–796 Represa A, Ben-Ari Y (2005) Trophic actions of GABA on neuronal development. Trends Neurosci 28:278–283 Roze E, Meijer L, Bakker A, van Braeckel KN, Sauer PJ (2009) Prenatal exposure to organohalogens, including brominated flame retardants, influences motor, cognitive, and behavioral performance at school age. Environ Health Perspect 117:1953–1958 Schantz SL, Widholm JJ, Rice DC (2003) Effects of PCB exposure on neuropsychological function in children. Environ Health Perspect 111:357–376 Seegal RF, Brosch KO, Okoniewski RJ (1997) Effects of in utero and lactational exposure of the laboratory rat to 2,4,2′,4′- and 3,4,3′,4′tetrachlorobiphenyl on dopamine function. Toxicol Appl Pharmacol 146:95–103 Shimokawa N, Miyazaki W, Iwasaki T, Koibuchi N (2006) Low-dose hydroxylated PCB induces c-Jun expression in PC12 cells. Neurotoxicology 27:176–183 Smaili SS, Pereira GJ, Costa MM, Rocha KK, Rodrigues L, do Carmo LG, Hirata H, Hsu YT (2013) The role of calcium stores in apoptosis and autophagy. Curr Mol Med 13:252–265 Spitzer NC (2006) Electrical activity in early neuronal development. Nature 444:707–712 Sugawara N, Nakai K, Nakamura T, Ohba T, Suzuki K, Kameo S, Satoh C, Satoh H (2006) Developmental and neurobehavioral effects of perinatal exposure to polychlorinated biphenyls in mice. Arch Toxicol 80:286–292 Tagliaferri S, Caglieri A, Goldoni M, Pinelli S, Alinovi R et al (2010) Low concentrations of the brominated flame retardants BDE-47 and BDE-99 induce synergistic oxidative stress-mediated neurotoxicity in human neuroblastoma cells. Toxicol In Vitro 24:116–122 Tilson HA, Kodavanti PR (1998) The neurotoxicity of polychlorinated biphenyls. Neurotoxicology 19:517–525

Environ Sci Pollut Res Tofighi R, Johansson C, Goldoni M, Ibrahim WN, Gogvadze V, Mutti A, Ceccatelli S (2011) Hippocampal neurons exposed to the environmental contaminants methylmercury and polychlorinated biphenyls undergo cell death via parallel activation of calpains and lysosomal proteases. Neurotox Res 19:183–194 Toms LM, Sjödin A, Harden F, Hobson P, Jones R et al (2009) Serum polybrominated diphenyl ether (PBDE) levels are higher in children (2–5 years of age) than in infants and adults. Environ Health Perspect 117:1461–1465 Ulbrich B, Stahlmann R (2004) Developmental toxicity of polychlorinated biphenyls (PCBs): a systematic review of experimental data. Arch Toxicol 78:252–268 Usenko CY, Hopkins DC, Trumble SJ, Bruce ED (2012) Hydroxylated PBDEs induce developmental arrest in zebrafish. Toxicol Appl Pharmacol 262:43–51 Van den Berg M, Birnbaum LS, Denison M, De Vito M, Farland W et al (2006) The 2005 World Health Organization reevaluation of human and mammalian toxic equivalency factors for dioxins and dioxin-like compounds. Toxicol Sci 93:223–241 Viberg H, Fredriksson A, Eriksson P (2003a) Neonatal exposure to polybrominated diphenyl ether (PBDE 153) disrupts spontaneous behaviour, impairs learning and memory, and decreases hippocampal cholinergic receptors in adult mice. Toxicol Appl Pharmacol 192:95–106 Viberg H, Fredriksson A, Jakobsson E, Orn U, Eriksson P (2003b) Neurobehavioral derangements in adult mice receiving decabrominated

diphenyl ether (PBDE 209) during a defined period of neonatal brain development. Toxicol Sci 76:112–120 Voie ØA, Fonnum F (1998) Ortho-substituted polychlorinated biphenyls elevate intracellular [Ca2+] in human granulocytes. Environ Toxicol Pharmacol 5:105–112 Westerink RHS (2006) Targeting exocytosis: ins and outs of the modulation of quantal dopamine release. CNS Neurol Disord Drug Targets 5:57–77 Westerink RHS, Vijverberg HPM (2002) Vesicular catecholamine release from rat PC12 cells on acute and subchronic exposure to polychlorinated biphenyls. Toxicol Appl Pharmacol 183:153–159 Winneke G, Walkowiak J, Lilienthal H (2002) PCB-induced neurodevelopmental toxicity in human infants and its potential mediation by endocrine dysfunction. Toxicology 181–182:161–165 Xing T, Chen L, Tao Y, Wang M, Chen J, Ruan DY (2009) Effects of decabrominated diphenyl ether (PBDE 209) exposure at different developmental periods on synaptic plasticity in the dentate gyrus of adult rats in vivo. Toxicol Sci 110:401–410 Xing TR, Yong W, Chen L, Tang ML, Wang M, Chen JT, Ruan DY (2010) Effects of decabrominated diphenyl ether (PBDE 209) on voltage-gated sodium channels in primary cultured rat hippocampal neurons. Environ Toxicol 25:400–408 Yu K, He Y, Yeung LWY, Lam PKS, Wu RSS, Zhou B (2008) DE-71induced apoptosis involving intracellular calcium and the Baxmitochondria-caspase protease pathway in human neuroblastoma cells in vitro. Toxicol Sci 104:341–351